You are on page 1of 15

APOR 363

Applied Ocean Research 21 (1999) 1–15

On the water impact of general two-dimensional sections


Xiaoming Mei, Yuming Liu, Dick K.P. Yue*
Department of Ocean Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA
Received 8 April 1998; accepted 1 October 1998

Abstract
We consider the analytic solution of the impact problem of a general two-dimensional body entering initially calm water. Of interest are
the water splash-up height, the force history and the pressure distribution on the body. The potential-flow formulation of Wagner (Wagner, H.
Math. Mech. 1932;12(4):193–215) is applied and extended to an arbitrary body section with the body boundary condition imposed on the
exact wetted surface of the body. For wedges and circular cylinders, we derive closed-form solutions using conformal mapping for the
boundary-value problem at any instant. These solutions reduce to those of Wagner in the small deadrise angle/shallow-body limit and are
verified for the general case by the existing experiments and fully nonlinear numerical simulations. For ship-like sections, we develop a
general scheme based on Lewis-form representations for which we also obtain analytic solutions. For illustration, the solution for a bow flare
section is obtained which compares favorably with experiments. The present approach generalizes Wagner’s method to a wide class of body
sections and is of immediate practical use in the study of ship slamming. 䉷 1999 Elsevier Science Ltd. All rights reserved.
Keywords: Impact; Slamming; Conformal-mapping; Lewis-form; Wedge

1. Introduction deadrise angle, it is implicitly given in terms of integral


equations, which must be solved numerically. Such similar-
The impact of bluff bodies striking a water surface often ity solutions, of course, cannot be obtained for arbitrary
causes an extremely large load with high transient local bodies.
pressure on the bodies. This slamming load can result in Recently, Zhao et al. [7] generalized the work of Wagner
substantial damage in the structures of ship hulls and marine [1] by applying linearized free-surface boundary conditions
facilities. Owing to its practical importance in ocean engi- on the horizontal plane at the splash-up height and imposing
neering, the water-body impact problem has attracted a the body boundary condition on the actual position of the
large number of investigations (for a review, see [2]). body. They solved the problem numerically using a bound-
An important pioneering work can be attributed to von ary-integral equation method. For simplicity, they ignored
Kármán [3] who developed an asymptotic theory for the flat the near-field flow at the water intersection. For wedges and
and near-flat impact problems with the linearized free a section with flare, their results agree well with experiments
surface and body boundary conditions. Later, Wagner [1] and fully nonlinear numerical simulations [8] for the slam-
modified von Kármán’s solution by accounting for the effect ming force and the body pressure distribution. This finding
of water splash-up on the body during the impact. This work suggests that the local jet flow at the water intersection does
was further developed by Armand and Cointe [4] and Howi- not play an appreciable role in the overall dynamics of the
son et al. [5], who included the effect of nonlinear jet flow in impact.
the intersection region between the body and free surface The present study is motivated by the work of Zhao et al.
using asymptotic matching. Owing to the inherent assump- [7], but is focused on the derivation of an analytic solution
tion originally in von Kármán [3], all these theories are for the general impact problem. We adopt their assumptions
applicable only to impact with shallow body submergence. and use the same initial-boundary-value formulation, but
For the special case of wedges entering water vertically at solve the boundary-value problem at any instant analytically
constant velocity, Dobrovol’skaya [6] derived an analytic using conformal mapping technique. Analytic solutions are
solution by taking advantage of the simplicity of the body obtained for wedges, circles, and general section geometries
geometry. Although this similarity solution is valid for any represented by Lewis forms [9,10]. These solutions compare
well with the existing experimental measurements. For
* Corresponding author. numerical confirmation, we develop a fully nonlinear
0141-1187/99/$ - see front matter 䉷 1999 Elsevier Science Ltd. All rights reserved.
PII: S0141-118 7(98)00034-0
2 X. Mei et al. / Applied Ocean Research 21 (1999) 1–15

Fig. 1. Definition sketch.

simulation for this problem using a Cauchy-integral method boundary condition takes the form
[11] with a matching jet solution near the water intersection
Dh 2F
point. The analytical and fully nonlinear numerical predic- ˆ on z ˆ h…y; t†; …2†
Dt 2z
tions are in excellent agreement in the case of a wedge (for
arbitrary deadrise angles). where D/Dt is the substantial derivative. Upon assuming no
This article is organized as follows. In Section 2, the air pocket to be involved in the impact and setting the atmo-
mathematical formulation and partial linearization of the spheric pressure to be zero, the dynamic free-surface bound-
general impact problem are presented. The analytic scheme ary condition can be expressed as:
to solve the resulting initial-boundary-value problem using
DF 1
conformal mapping is given in Section 3. In Section 4, we ⫺ …F2y ⫹ F2z † ⫹ gh ˆ 0 on z ˆ h…y; t†; …3†
Dt 2
derive the solution for wedges and make comparisons to the
existing experiments and results of a fully nonlinear numer- where g is the gravitational acceleration. On the body, no-
ical method. The solution for circular sections and its flux condition is imposed:
comparison to experiments are given in Section 5. For 2F
general section geometries, we develop an analytic proce- ˆ ⫺Vnz on SB …t†; …4†
2n
dure based on the use of Lewis-form representations in
Section 6. As an illustration, the solution for a bow-flare where n ˆ …ny ; nz † is the unit normal out of the body. In Eq.
section is obtained and compared with the measurements. (4), SB …t† represents the instantaneous wetted body surface
Conclusions are presented in Section 7. which is determined by h ˆ H…y† with y ⱕ jY…t†j; where
Y(t) denotes the horizontal coordinate of the intersection
between the body and free surface (see Fig. 1). For initial
2. Problem formulation conditions, at the moment of initial contact, t ˆ 0, we have:
F…y; z; 0† ˆ h…y; 0† ˆ 0 on z ˆ 0: …5†
We consider the two-dimensional water-entry problem
involving a (rigid) body forced through an initially calm The initial-boundary-value problem for F is complete with
(horizontal) water surface at a constant vertical velocity V. the imposition of the far-field radiation condition:
For simplicity, the cross-section of the body is assumed to
be symmetric about its vertical centerline. We introduce a j7Fj ! 0 as y2 ⫹ z2 ! ∞: …6†
Cartesian coordinate system (y, z) with the y-axis in the In terms of the velocity potential F, the pressure on the
undisturbed water plane and z-axis positive upward coincid- body is determined according to Bernoulli equation, and the
ing with the vertical centerline of the body (see Fig. 1). We impact force on the body can be obtained by direct integra-
represent the body surface by the relation h ˆ H…y†, where tion of the pressure over the wetted body surface.
h denotes the vertical distance of a point on the body We note that the stated problem is nonlinear owing to the
measured from the (bottom) apex of the body. free-surface boundary conditions (2) and (3). In this case,
We assume the fluid to be incompressible and inviscid the problem must be solved numerically even for simple H.
and the fluid motion irrotational. The flow at any time t can To understand the mechanics of the impact, we seek analy-
be described by a velocity potential F…y; z; t† which satisfies tic solutions to this problem. To do this, we consider a
the Laplace equation in the fluid domain partial linearization of the problem along the lines of
72 F ˆ 0: …1† Wagner [1] and Zhao et al. [7].
If we consider the characteristic time to be the duration of
On the free surface, denoted by z ˆ h…y; t†, the kinematic the impact, T; the characteristic free-surface wavelength the
X. Mei et al. / Applied Ocean Research 21 (1999) 1–15 3

Fig. 2. Relationship between the (y, z) and …y 0 ; z 0 † coordinate systems.

size of the body cross-section, L; and the characteristic wave We note that the position of the intersection Y(t) between the
amplitude the water-entry depth of the body, VT, the char- body and free surface is a priori unknown, which must be
acteristic velocity potential must be VL and the free-surface solved together with the initial-boundary-value problem.
boundary conditions (2) and (3) can be scaled accordingly. The approximation discussed above is invalid for the
Denoting scaled variables by ( ˜ ); Eqs. (2) and (3) become: local flow near the water intersection where the free surface
2h~ rises sharply and full nonlinearity must, in principle, be
⫹ 1h~ y~ F~ y~ ˆ F~ z~ on z~ ˆ 1h~ ; …7† retained (see, e.g. [1,4]). For the global impact force and
2t~
the overall pressure distribution. However, Zhao et al. [7]
and showed that the effect of the local jet flow is quite small. In
 
1  ~ 2  ~ 2
the present study, we ignore the effect of the jet flow. Our
DF~ 12
⫺ F y~ ⫹ F z~ ⫹ 2 h~ ˆ 0 on z~ ˆ 1h~ ; solution can be considered to be the outer-field solution of
Dt~ 2 Fr the overall problem.
…8†
where 1 ⬅ VT=L measures the surface wave steepness, and 3. Analytic solution scheme
Fr ⬅ V=…gL†1=2 is the Froude number.
If the free-surface wave slope is small, i.e., 1 ˆ VT=L p In this section, we outline an analytic scheme for the
1; the nonlinear terms in Eqs. (7) and (8) can be neglected. general solution of the simplified impact problem. Unlike
Comparable to the nonlinear terms, the last term in Eq. (8) Wagner [1], where the body boundary condition is applied
can be dropped provided that the Froude number is not on the projection of the body on the free surface, we impose
small, specifically, if Fr ⱖ O…11=2 †: This implies that the the body boundary condition at the instantaneous position of
entry velocity V must be O(gT) or larger. Assuming this to the body (see [4]). The present results are thus not restricted
be the case, we obtain the simplified dynamic free-surface to shallow submerged body shapes or small body deadrise
boundary condition. angles.
F~ …y;
The simplified initial-boundary-value problem is solved
~ z~; t~† ˆ 0 on z~ ˆ 1h~ : …9†
in two steps, both of which can be obtained analytically: (I)
Strictly speaking, as 1 p 1, Eq. (9) can be further linearized for a given intersection position Y(t), obtain the vertical
to apply on z~ ˆ 0: Following Wagner [1], to account for the velocity of the free surface from the boundary-value
splash-up effect, we write as problem; and (II) integrate the evolution Eq. (12) in time
to find the new intersection position.
F~ …y;
~ z~; t~† ˆ 0 on z~ ˆ 1h~ …Y; ~ Y†
~ t~† ˆ 1…H… ~ ⫺ t~ †: …10†
The specification of F~ ˆ 0 on z~ ˆ 1h~ …Y;~ t~ † rather than on 3.1. Analytic solution of the boundary-value problem
z ˆ 0 is the key difference between the theory of von The boundary-value problem for F consists of the
Kármán [3] and that of Wagner [1]. Laplace equation (1), the Dirichlet free-surface boundary
Returning to physical variables, the simplified initial- condition Eq. (11), the Neumann body boundary condition
boundary-value problem for F consists of the Lapalace Eq. (4), and the radiation condition Eq. (6).
Eq. (1), the initial condition Eq. (5), the radiation condition The condition F ˆ 0, applied on the plane z ˆ h…Y†,
Eq. (6), the body boundary condition (4), and the free- implies that F is antisymmetric about the plane z ˆ h…Y†:
surface dynamic boundary condition As a result, F can be considered to be a closed ‘‘double’’
F…y; z; t† ˆ 0 on z ˆ h…Y; t† ˆ H…Y† ⫺ Vt: …11† body moving in an unbounded fluid with a constant down-
ward velocity V. This fictitious ‘‘double’’ body, DB (say),
The simplified free-surface kinematic boundary condition consists of the immersed part of the body, SB …t†; and its
takes the form image about the plane z ˆ h…Y†. We now consider a coor-
2h dinate system (y 0 , z 0 ) fixed on DB with the y 0 - and z 0 - axes
ˆ Fz on z ˆ H…Y† ⫺ Vt: …12† coinciding with the horizontal and vertical symmetry lines
2t
4 X. Mei et al. / Applied Ocean Research 21 (1999) 1–15

of DB (see Fig. 2). The (y 0 , z 0 ) coordinate system is related to complicated and a closed-form solution for m…`† from (16)
(y, z) by the Galilean transformation: y 0 ˆ y and z 0 ˆ z ⫺ cannot be obtained. For bodies with smooth surfaces, the
h…Y† with intersection point moves smoothly with time and hence
m…`† varies smoothly with `: In such cases, we can write
F…y; z; t† ˆ f…y 0 ; z 0 ; t† ⫺ Vz 0 ; …13†
an expansion for m…`†; say, in terms of Chebyshev polyno-
where f represents the velocity potential for the ‘‘double- mials as
body’’ problem. X
∞  
To solve for f , we apply a conformal mapping with the m…`† ˆ an Tn …`† for ` 僆 0; Ymax ; …17†
body boundary condition satisfied exactly. In general, the nˆ0
mapping depends on the body geometry, and hence the where Tn …`† ⬅ Tn …2`=Ymax ⫺ 1†; and Tn is the nth order
solution f . As DB depends on Y(t), the velocity potential Chebyshev polynomial of the first kind. In (17), an are
f is ultimately a function of the intersection position Y(t). unknown coefficients, and Ymax is the maximum value of Y
For relatively simple SB, such as wedges and circles, the reached during the
mapping and hence the complete solution can be obtained   impact. For a sufficiently smooth
m…`†; ` 僆 0; Ymax , (17) converges exponentially. In prac-
in closed form. These are presented in Sections 4 and 5, tice, we truncate the series at some n ˆ N for a specified
respectively. For ship-like sections representable in terms accuracy. To solve ` ˆ Y…t†; it is convenient to write (17)
of Lewis forms, analytic solutions can likewise be obtained as
as given in Section 6.
X
N X
N  
3.2. Evolution of the position of the intersection point m…`† ˆ an Tn …`† ˆ bn ` n for ` 僆 0; Ymax ;
nˆ0 nˆ0
The determination of the intersection point I ˆ …18†
…Y…t†; H…Y† ⫺ Vt† is essential for solving the overall
where
problem to obtain quantities such as impact force and pres-
sure. Following Wagner [1], we define I to be the intersec- X
n
bn ˆ ak ckn ; n ˆ 0; 1; …; N; …19†
tion of the rising free surface h…y; t† and the body surface
kˆ0
SB …t† : H…y† ⫺ Vt: With this definition and by applying the
kinematic free-surface boundary condition (12), we obtain and ckn are the known constants given by the definition of
the equation governing the position of the intersection point Chebyshev polynomials.
with Y(t) satisfying From Eq. (18), the time dependence of the intersection
Zt position Y(t) can be obtained as follows:
H…Y† ⫺ Vt ˆ v…Y; t† dt; …14† X
N
0 bn
Vt ˆ Y n⫹1 : …20†
where t is the dummy time variable and v…Y; t† ˆ ht …Y; t† is nˆ0
n⫹1
the vertical velocity of the free surface at y ˆ Y…t†: From the
To determine the unknown coefficients an ; n ˆ 0; 1; …; N;
boundary-value solution, v…Y…t†; t† is given by
we substitute Eq. (18) into Eq. (16) to obtain
2 2
v…Y…t†; t† ⬅ F…Y…t†; h…Y…t††; t† ˆ 0 f…Y…t†; 0; t† ⫺ V: X
N  
2z 2z H…Y† ˆ an bn …Y† for Y 僆 0; Ymax ; …21†
…15† nˆ0

As v…Y…t†; t† depends on the history of the intersection point, where the influence coefficient bn ; n ˆ 0; 1; …; N; is given
y ˆ Y…t†; t ⬍ t; which itself is a priori unknown, (14) with by
(15) provide the implicit time dependence for the intersec- ZY
tion point position Y(t). bn …Y† ˆ v0 …Y; `†Tn …`† d`; n ˆ 0; 1; …; N: …22†
0
To derive the direct time dependence, we substitute (15)
into (14), replace the integral with respect to t by one with For a wedge, Eq. (21) can be solved analytically (see
respect to ` ⬅ Y…t† (i.e., interchanging the dependent and Section 4.1) but for more general geometries, (21) must
independent variables), and obtain be solved numerically. To do that, we apply Eq. (21) at
ZY…t† N ⫹ 1 discrete points
 on the body surface corresponding
H…Y…t†† ˆ v0 …Y…t†; t…`††m…`† d`; …16† to the range Y 僆 0; Ymax , and solve the resulting linear
0
system of equations for an ; n ˆ 0; 1; …; N: For exponential
where m…`† ⬅ V…dt=d`† and v0 ⬅ v=V: Eq. (16) is an inte- convergence with Chebyshev polynomials, the collocation
gral equation for the unknown m…`† and can, in principle, be points should be at the maxima of TN : We note that for
solved provided the dependence of the kernel v0 …Y…t†; t…`†† typical (smooth) geometries, N can be quite small in prac-
on ` is known. tice, e.g., N ˆ O(10) for 4D accuracy for a circular section
For arbitrary bodies, the dependence of v0 on ` ˆ Y…t† is (see Section 5, Table 1).
X. Mei et al. / Applied Ocean Research 21 (1999) 1–15 5

Table 1
Convergence with increasing the order N of the coefficients, an,n ˆ 0,1,…,N of the Chebyshev polynomial expansion for the velocity of the intersection point in
the water entry of a circular cylinder. Ymax/R ˆ 0.96

N a0 a1 a2 a3 a4 a5 a6 a7 a8 a9 a10

4 0.4908 0.6520 0.2309 0.0975 0.0280


5 0.4975 0.6648 0.2499 0.1164 0.0475 0.0139
6 0.5016 0.6747 0.2603 0.1299 0.0599 0.0242 0.0071
7 0.5048 0.6808 0.2680 0.1379 0.0689 0.0320 0.0125 0.0037
8 0.5068 0.6855 0.2728 0.1437 0.0746 0.0377 0.0173 0.0065 0.0019
9 0.5114 0.6938 0.2804 0.1502 0.0808 0.0430 0.0216 0.0098 0.0036 0.0010
10 0.5096 0.6912 0.2792 0.1507 0.0822 0.0452 0.0241 0.0120 0.0054 0.0020 0.0006

3.3. Impact pressure and force on the body physical experiments, and fully nonlinear numerical simu-
lations.
After obtaining the time dependence of the intersection
position Y(t), the potential F for any specified Y(t) is given 4.1. Analytic solution
by the (conformal mapping) solution of the ‘‘double-body’’
boundary-value problem. In term of the velocity potential For the wedge, the ‘‘double-body’’ velocity potential f
F, the pressure on a point on the body is given by Bernoulli describes a uniform vertical flow past a rhombus with width
equation 2Y(t) and entry angle p ⫺ 2a. The solution is obtained in a
closed form by mapping the flow in the physical plane,
P…y; t† 2F 1 Z ˆ y 0 ⫹ iz 0 , into a uniform vertical stream in the mapped
ˆ⫺ ⫺ …F2y ⫹ F2z †
r 2t 2 plane, W ˆ p ⫹ iq, using Schwartz–Christoffel transforma-
tion. The solution can be written in the form
DF 1
ˆ⫺ ⫺ V Fz ⫺ …F2y ⫹ F2z †; …23†
Dt 2 f…y 0 ; z 0 ; t† ⬅ f…Z; t† ˆ Re{ ⫺ iW}VY…t†=A; …26†
where r is the fluid density when the pressure at infinity is where Z and W are related by the conformal mapping
set to be zero, and gravity effect is ignored. On using Eq. !u
ZW w2
(13) with z 0 ˆ H…y†⫺H…Y†, Eq. (23) can be rewritten in Z=Y…t† ˆ A ⫺1
dw ⫹ 1: …27†
terms of f as 0 w2 ⫹ 1
P…y; t† Df 1 D 1 Eq. (27) maps the ‘‘double-body’’ rhombus in the Z plane
ˆ⫺ ⫺ …f2y 0 ⫹ f2z 0 † ⫺ V H…Y† ⫹ V 2 :
r Dt 2 Dt 2 to the vertical line segment from (0, ⫺ 1) to (0, 1) in the W
…24† plane. In this case, the coefficient A is a function of a and is
given by
As in Wagner [1] and Zhao and Faltinsen [8], the nonlinear !u
terms in Eqs. (23) and (24) are included to offset the singular Z1 w2
A…a† ˆ cos a dw with u ˆ …p ⫺ 2a†=2p:
0 1⫺w
effect at the intersection point from the linear term DF/Dt. 2

The (vertical) force on the body, F(t), can be obtained by


…28†
direct integration of the pressure over the wetted body
surface. This force can also be determined using the defini- The velocity of the flow is obtained from Eq. (26) as
tion of added mass and conservation of vertical momentum. !u
⫺1 dW W2 ⫹ 1
d u ⫺ iv ˆ ⫺iVY…t†A ˆ ⫺iV : …29†
F…t† ˆ …M V†; …25† dZ W2
dt a
Knowing f at any instant, the motion of the intersection
where Ma is the infinite-frequency added mass in heave for point Y(t) can be determined according to Eq. (20) via the
that protion of the body below H(Y(t)). coefficients bn ; n ˆ 0; 1; …; N; which are given by Eqs. (21)
and (19). To evaluate the functions bn ; n ˆ 0; 1; …; N; we
first determine the vertical velocity at Z ˆ Y…t† and time
4. Application to wedges t; t ⱕ t; from Eq. (29).
!u
We present the analytic solution for the water impact of a p2 ⫹ 1
two-dimensional wedge described by h ˆ H…y† ˆ y tan a v0 …Y; `…t†† ˆ ; …30†
p2
where a is the deadrise angle. The present result is a gener-
alization of Wagner [1], valid for small deadrise angles to where W ˆ p is the image in the W-plane of the physical
arbitrary deadrise angle. We verify the analytic solution by point Z ˆ Y related by the mapping (27). Upon substituting
making comparisons with the existing asymptotic theories, Eq. (30) into Eq. (22) and using Eq. (27), we can easily show
6 X. Mei et al. / Applied Ocean Research 21 (1999) 1–15

Fig. 3. The relative water splash-up, g , as a function of the deadrise angle, a , of a wedge entering vertically into water at constant velocity. Results plotted are:
– · –, Pierson’s hypothesis [12]; —, the present exact solution (36); and – – –, the Lewis-form approximation solution (65).

that which is the solution of von Kármán [3] in the limit of no


X
n splash-up. Fig. 3 also shows the result from the so-called
bn …y† ˆ ckn fk …a†Y k⫹1 ; n ˆ 0; 1; …; N; …31† Pierson’s hypothesis [12], which assumes g to vary linearly
kˆ0 with a between the two above limits. Except at the two
where the constant fk …a† is given by limits of a ˆ 0 and p /2. Pierson’s hypothesis considerably
! underestimates the splash-up coefficient, with g ⫺ 1 less
Z 1 p2 ⫹ 1 u than half of the analytic solution (36) for a ⬎t p=3.
f k … a† ˆ lk d l k ˆ 0; 1; …; N: …32† From Y(t), the velocity potential f can be evaluated from
0 p2
Eqs. (26) and (27) at any instant t and the problem is solved
Here l ⬅ `=Y and p is a function of l given by the relation completely. The pressure on the body can be obtained from
!u Eq. (24) and the pressure coefficient Cp …y; t† ⬅
Zp w2
⫺1
…l ⫺ 1†A ˆ dw: …33† P…y; t†=…rV 2 =2†; y ⱕ Y…t†; is given by
0 w ⫹1
2
" !u !u #
2g 1 ⫺ q2 Z 1 w2
As H…Y† ˆ Y tan a for the wedge, it follows from (21) that Cp ˆ dw ⫺ q
A tan a q2 jqj 1 ⫺ w
2
a0 ˆ tan a=f0 and an ˆ 0 for n ˆ 1; …; N: It follows from
(19) that b0 ˆ a0 and bn ˆ 0 for n ⬍ 0. Finally from (20) we !2u
1 ⫺ q2
have ⫺ ⫺2g ⫹ 1; …37†
q2
Y…t† ˆ Vtf0 …a†=tan a: …34†
where the real variable q is related to the position on the
The motion of the intersection between the body and the
body by
free surface is often described in terms of the vertical coor-
!u
dinate h…t† ˆ Y…t† and tan a . It follows that h…y† Zq w2
⫺1
ˆ gA cos a dw ⫹ g
h…t† ˆ Vtf0 …a† ˆ Vtg…a†; …35† Vt 0 1⫺W
2
…38†
where g ⫺ 1 is the splash-up coefficient which measures the h…y†
rise of the water surface on the body relative to its submer- for 0 ⱕ ⱕ g:
Vt
gence Vt. Clearly, g is a function of a and takes the form
(cf. Eqs. (32) and (33)): Near the intersection, Y(t) ⫺ y p 1, the solution (37)
" !u #⫺2 behaves asymptotically as:
Z∞ Zp w2
g…a† ⬅ f0 …a† ˆ A dw ⫹ A dp: …36† 2g
0 w ⫹1
2 Cp ˆ jqj⫺4u ⫹ ⫺2u
jqj ⫹O…1†
0
sin a …39†
Fig. 3 plots the solution of g as a function of a . For  1=…2u⫹1†
for jqj ˆ A…2u ⫹ 1†…1 ⫺ y=Y†=cos a p 1:
a ˆ 0, we have g ˆ p=2 which agrees with the asymptotic
solution of Wagner [1] for a p 1. For a ˆ p=2, g ˆ 1 Clearly, the pressure is singular at the intersection point,
X. Mei et al. / Applied Ocean Research 21 (1999) 1–15 7

Fig. 4. The pressure distributions on the wedge and the associated free-surface profiles near the free-surface intersection region at different deadrise angles a .
The results plotted are: —, the present analytic solution; and – – –, the similarity solution [6].
8 X. Mei et al. / Applied Ocean Research 21 (1999) 1–15

Fig. 5. Comparisons of the force coefficient as a function of the deadrise angle a for a wedge entering vertically into water at constant velocity. The results
plotted are: —, the present solution by direct pressure integration; – – –, the present solution using the added-mass formula; – · –, the asymptotic solution of
van Kármán (1932); – · · –, the asymptotic solution of Wagner [1]; O, the similarity solution [6]; and …, the approximate solution of Payne [12] applying
Pierson’s hypothesis.

y ˆ Y, but the resulting impact force is finite as the singu- Zhao et al. [7] within the graphical error and thus their
larity is integrable. results are not plotted.
The instantaneous free-surface shape is obtained by inte- The slamming force on the body can be obtained by direct
grating the kinematic free-surface boundary condition (12) integration of the pressure distribution or with the use of Eq.
with time. The result can be expressed in terms of a simi- (25). Fig. 5 compares the impact force coefficient, as a func-
larity variable x ⬅ y=Vt as tion of the deadrise angle, obtained using these two calcula-
tions. For the added-mass theory, the added mass of the
h…y; t† h … x†
ˆ wedge is calculated by the commonly used form
Vt Vt
" !u #⫺2 Ma ˆ Ca rY 2 ; Ca ˆ …1 ⫺ a=2p†2 p=2: …42†
x tan a Z∞ ⫺1 Zp w2
ˆ A dw ⫹ 1 dp ⫺ 1;
Ag p0 0 1⫹w
2
From Fig. 5 the added-mass theory result is not equal to, but
…40† is somewhat larger than that given by pressure integration.
where the lower limit of the integral p0 depends on x and is This is expected owing to the nonlinear terms in the
determined from: Bernoulli equation (Eq. (24)), which gave a negative contri-
!u bution to the impact force obtained with direct pressure
Zp0 w2 integration, while Eq. (42) is equivalent to that obtained
⫺1
x tan a ˆ gA dw ⫹ g: …41† using the linearized Bernoulli pressure [13]. The added-
0 1 ⫹ w2
mass result should thus be considered to be an approxima-
Fig. 4 shows the pressure distributions on the body and the tion.
associated free-surface profiles for a range of deadrise Also shown for comparison in Fig. 5 are force coefficients
angles a . The negative pressure caused by the singularity obtained from the theories of von Kármán [3], and Wagner
at the intersection (cf. (39)) is omitted in Fig. 4, as it is not [1]; the similarity solution of Dobrovol’skaya [6]; and the
physical and does not exist if the local intersection solution hypothesis of Pierson [12]. As expected, the asymptotic
is included (cf. [8]). From Fig. 4, we observe that, for small theory of von Kármán [3] underestimates the slamming
a , the maximum pressure is obtained near the intersection force, while that of Wagner [1] overestimates it. However,
between the body and free surface, while for large a , the the similarity result agrees well with the present theory. The
maximum is located at the keel point. For comparison, the solution of Payne [12] is derived using the Pierson’s hypoth-
similarity solution [6] is also shown in Fig. 4. The present esis and the added-mass equation (Eq. (25)). Remarkably,
theory agrees well with the similarity solution. We remark despite the poor prediction of the splash-up (see Fig. 3), this
that the present solution agrees with the numerical result of solution agrees well with the present (pressure-integration)
X. Mei et al. / Applied Ocean Research 21 (1999) 1–15 9

Fig. 6. Comparisons of the impact force on a freely falling wedge, a ˆ 30⬚, as a function of time t. The results plotted are: – – –, present theory; —,
experimental measurement of Zhao, Faltinsen and Aarsnes [7]; and the theories of: – · · –, von Kármán [3]; and – · –, Wagner [1].

result. This is owing to the fact that the effect of under- and Eq. (44) becomes
estimation of splash-up by the Pierson’s hypothesis just
rV02 Y
happens to offset the impact force overestimation as a result F…t† ˆ 2Ca gtan⫺1 a : …46†
of using the added-mass formula. …1 ⫹ Ma =M†3
Using Eqs. (43) and (45), the motion of the intersection
4.2. Comparison to experiments during water entry can be determined upon eliminating V
and is given by
We compare our analytic solution to the experiment of
Zhao et al. [7] who examined the free falling of a wedge into rC a 3 g V0 t
Y ⫹Y ˆ : …47†
calm water. The model has a deadrise angle of 30⬚. As the 3M tan a
velocity of the body in the free-fall experiment varies with The mass of the model is M ˆ 241 kg=m in the experi-
time after water entry, the analytic results of Section 4.1 for ment; the aspect ratio of the model is about 5, and we follow
constant falling velocity must be modified for direct Zhao et al. [7] and multiply the added mass in Eq. (42) by a
comparisons. Upon applying momentum theorem and factor of 0.8 to account for three-dimensional effects in the
ignoring gravity, the instantaneous velocity of the wedge experiment. For a ˆ 30⬚, our analytic solution (36) gives
during impact is given by g ˆ 1.51. With these values in Eq. (47) to obtain Y(t) and
V0 substituting into Eq. (44), we obtain a theoretical prediction
V…t† ˆ ; …43† of the impact force for comparison to the measurement of
1 ⫹ Ma =M
Zhao et al. [7]. This is shown in Fig. 6. The comparison is
where V0 is the falling velocity of the body at initial impact, excellent until later time when flow separation on the body
M the mass of the body, and Ma the infinite-frequency added is observed and reported.
mass in heave for the submerged portion of the body, given Also plotted in Fig. 6 are the predictions using the
by Eq. (42). Note that the added mass here depends on the theories of von Kármán [3] (g ˆ 1) and Wagner [1] (g ˆ
position of the intersection Y(t). p /2). As expected, von Kármán’s solution significantly
From Newton’s second law, the impact force on the body underestimates the impact force while Wagner’s solution
is overestimates it.
 
dV 2Ca rYV0 dY
F…t† ⬅ ⫺M ˆ : …44† 4.3. Comparison to fully nonlinear numerical simulations
dt …1 ⫹ Ma =M†2 dt
Finally, we make comparisons to fully nonlinear numer-
Upon using the solution (34), we have
ical simulation, which includes the effect of the spray jet in
dY=dt ˆ V g=tan a; …45† the intersection region. The numerical scheme we use is an
10 X. Mei et al. / Applied Ocean Research 21 (1999) 1–15

Fig. 7. Evolution of the free surface profile caused by the water entry of a wedge with deadrise angle a ˆ 30⬚ and velocity V ˆ 1. Plotted are the free surface
elevations (increasing upwards) at times t ˆ 0.0, 0.001, 0.01, 0.02, 0.03, 0.04 and 0.05.

extension of the Cauchy-integral mixed-Eulerian–Lagran- includes a local analytic representation of the jet in the
gian method of Vinje and Brevig [11] but with a special global solution using a matching scheme.
matching treatment at the intersection point. In practice, We introduce the complex potential
the spray jet has large velocity but small thickness, which b…Z; t† ⬅ F…Z; t† ⫹ iC…Z; t†;where c is the associated
cause numerical simulations to break down (e.g. [14,15]). stream function. For the spray jet, we expand b in a Taylor
To overcome this difficulty, Zhao and Faltinsen [8] ignored series around the intersection point, Z ˆ ZI
the effect of the jet by simply cutting off the jet from the
global flow in the simulation. A drawback of this approach X
M

is that the simulation no longer conserves mass and momen- b…Z; t† ˆ b…ZI ; t† ⫹ Am …t†…Z ⫺ ZI †m : …48†
mˆ1
tum. Here, we propose an alternative approach, which

Fig. 8. Comparison of the pressure distribution on a wedge with deadrise angle a ˆ 30⬚ and constant dropping velocity V ˆ 1 between: – · –, analytic theory;
and O, fully nonlinear numerical simulation.
X. Mei et al. / Applied Ocean Research 21 (1999) 1–15 11

point. Except the thin jet, which is not included in the analy-
tic solution, the agreement between theory and nonlinear
simulation for both the free surface profile and the pressure
variation is excellent. Note that the analytic theory correctly
obtains the location of the maximum impact pressure, which
occurs near the root of the jet (where the free surface has a
maximum curvature) for this deadrise angle.

5. Application to circular cylinders

The analysis of Section 3 can be applied to obtain the


solution for the water entry of a circular section. Let the
radius of the circle be R. In the following, we normalize
Fig. 9. Comparisons of the free surface elevation and the impact pressure all lengths by this radius and set R ˆ 1. The circle below
for a wedge with deadrise angle a ˆ 30⬚ and constant dropping velocity
V ˆ 1 between: – · –, analytic theory; and O, fully nonlinear numerical
its center is given by h…y† ˆ 1 ⫺ …1 ⫺ y2 †1=2 : In this case, the
simulation. ‘‘double-body’’ potential f…y 0 ; z 0 ; t† at any instant t
describes the vertical uniform flow past a (doubly convex)
lens of width 2Y(t) and thickness 2h(Y) comprising the
In (48), the potential at the intersection point b…ZI ; t† is
immersed (wetted) portion of the circle and its image
generally known from the outer global solution obtained
about the z 0 ˆ 0 plane. Like the case of the wedge, a
from the body boundary condition and time integration of
closed-form solution for f can be obtained using conformal
the free-surface boundary conditions (see e.g. [14,16]). The
mapping.
coefficients Am ; m ˆ 1; 2; …; M are unknown, and are
We map the physical flow in the Z-plane into a uniform
obtained upon matching (48) to the outer numerical
vertical stream in the W-plane through a double conformal
solution.
mapping
Note that, as b is analytic in Z inside the fluid, the series
in (48) is convergent. In practice, the integrands of the iY iYv
Zˆ⫺ and W ˆ⫺ ; …49†
boundary integrals are represented by piecewise (unknown) tan Q sin…vQ†
functions outside the jet and by (48) within a small matching
where Q ˆ j ⫹ iz is an intermediate complex variable and
distance from the intersection point. The value of M
the dimensionless coefficient v is defined by
depends on the location of matching and the desired accu-
racy. An advantage of (48) is that the velocity of the jet can p=2
vˆ : …50†
be obtained directly by differentiation. arctan{Y=‰1 ⫺ …1 ⫺ Y 2 †1=2 Š}
The numerical results are checked for convergence by
In terms of the mapping variables W and Q, f can be written
varying systematically the time step, the discretization
in the form:
size of the body and free surface, the number of terms M
for the jet solution, and the matching distance from the f…y 0 ; z 0 ; t† ˆ Re{ ⫺ iVW}; …51†
intersection point. In addition, the global conservations of
and the complex velocity is
mass, momentum, and energy are verified. The maximum
numerical error for the results presented below is within 1%. v2 sin2 Q
Fig. 7 shows the fully nonlinear simulation result for the u ⫺ iv ˆ ⫺ iV : …52†
sin…vQ† tan…vQ†
evolution of the free-surface profile for a wedge with a ˆ
30⬚ and V ˆ 1. It is evident that the present matching treat- After obtaining the boundary-value solution in terms of
ment of the local spray jet at the free-surface intersection is the intersection position Y(t), the motion of the intersection
quite effective. For this wedge geometry, a value of M ˆ 2 point can be solved from Eq. (20). The vertical velocity on
in Eq. (48) is used. the free surface …z 0 ˆ 0† at y 0 ˆ Y…t† and time t; t ⱕ t; is
The comparison between the analytic solution and the determined from Eq. (52) to be:
fully nonlinear simulation result for the pressure distribution v2 sinh2 z
on the body is shown in Fig. 8. For both the location and the v0 …Y; `…t†† ˆ ; …53†
sinh…vz†tanh…vz†
magnitude of the maximum impact pressure, the analytic
theory agrees well with fully nonlinear simulation. The where z is real and related to ` by
simulation result also confirms that the dynamic pressure  
1 Y…t†=`…t† ⫺ 1
because of the jet flow is indeed relatively small. z ˆ z…t† ⬅ 1n : …54†
2 Y…t†=`…t† ⫹ 1
Fig. 9 shows the comparison between the analytic predic-
tion and the nonlinear simulation for the free surface profile To determine the unknown coefficients an from Eq. (21),
and the pressure distribution on the body near the intersection the coefficients bn ; n ˆ 0; 1; …; N; are evaluated by
12 X. Mei et al. / Applied Ocean Research 21 (1999) 1–15

Fig. 10. Evolution of the intersection point position in the water entry of a circular cylinder, radius R and velocity V. The results plotted are: —, the present
theory; – · –, solution using Wagner [1]; and – – –, the Lewis-form approximation solution of Section 6.

substituting Eq. (53) into Eq. (22) to obtain Table 1 gives values of numerically computed coeffi-
cients an ; n ˆ 0; 1; …; N for Ymax =R ˆ 0:96 for a range
ZY v2 sinh2 zT …`† of N. The rapid convergence of an with N is evident.
bn …Y† ˆ n
d` for n ˆ 0; 1; …; N:
0 sinh…vz† tanh…vz† For N ⱖ 10; an has converged to four decimal places.
…55† Such rapid convergence is obtained for all Ymax/R not
very close to 1. The latter is a singular point which
As the dependences of v and z upon ` are involved (cf. Eqs. corresponds to a physical separation of the flow at
(50) and (54)) the integral in Eq. (55) cannot be evaluated in Y ˆ R. Note that the converged row of values, say for
closed form. In this case Eq. (21) is solved numerically for a N ˆ 10, completely specifies the solution to the stated
given value of Ymax/R as explained in Section 3.2. problem.

Fig. 11. Impact force on a circular cylinder entering water with constant velocity. The results plotted are: present theoretical predictions using: —, direct
pressure integration; and –-, added-mass formula; – –, experiments of Campbell and Weynberg [17]; experiments of Armand and Cointe [4] with dropping
velocity: – · · –, V ˆ 7:38 m=s; and –·–, V ˆ 2:33 m=s; and the asymptotic theories of: …, von Kármán [3]; and …, Wagner [1].
X. Mei et al. / Applied Ocean Research 21 (1999) 1–15 13

We remark that the present theory recovers the result of wetted body surface by a Lewis form:
Wagner [1] for lesser time. In this case an asymptotic solu-
X
3
tion for (55) can be derived. For Vt p 1, we have Y p 1, Z0 …t† ˆ Y ⫹ ih ˆ H…t† An ei…3⫺2n†u for p ⱕ u ⱕ 2p;
and the cylinder surface can be represented by h…Y† ˆ nˆ1
Y 2 =2 ⫹ O…Y 4 †: From Eq. (50), we have v ˆ 1 ⫹ O(Y), …58†
which leads to v0 …Y; `† ˆ Y=…Y 2 ⫺ `2 †1=2 ⫹ O…Y† from
(53). From Eq. (22), we obtain where H…t† ˆ H…Y…t†† is the immersed draft of the body.
The dimensionless coefficients, A1 ; A2 and A3 , are defined
Xn Z1
by
bn …Y† ˆ Y k⫹1 Ckn lk =…1 ⫺ l2 †1=2 dl‰1 ⫹ O…Y†Š: …56†
0
kˆ0 1
A1 ˆ …L ⫹ 1† ⫺ A3 ; …59†
Solving Eq. (21), we have a0 ˆ a1 ˆ 1=4 and an ˆ 0 for n 2
⬎ 1. It follows from (19) that b1 ˆ a0 ⫹ a1 ˆ 1=2; and Y ˆ
1
2…Vt†1=2 ⫹ O…Vt† for Vt p 1 from Eq. (20). This solution is A2 ˆ …L ⫺ 1†; …60†
identical to that obtained with Wagner’s theory [13]. 2
Fig. 10 shows that position of the intersection point   
1 1 4s 1=2
obtained using Table 1 (with N ˆ 10). For comparison, A3 ˆ ⫺ …L ⫹ 1† ⫹ …L ⫹ 1† ⫹ 8L 1 ⫺
2
;
Wagner’s solution corresponding to Eq. (56) is also plotted. 4 4 p
As expected, the two solutions agree well for lesser time but …61†
differ significantly beyond the initial stage of impact. In where L ⬅ Y=H; s ⬅ S=2YH; and S is the area of the
particular, Wagner’s solution considerably overestimates immersed body section. As the intersection position Y varies
the speed of the intersection point motion. with time, in general, these coefficients are functions of time
Finally, we make comparison to the experiments of t. With this approximation for the body surface, the physical
Campbell and Weynberg [17] and Armand and Cointe [4] flow in the Z-plane can now be mapped into a uniform
for the impact force on the cylinder. In these experiments, vertical steam past a unit circle in the W-plane using the
the cylinder is forced into the water with constant velocities. following mapping:
For the theoretical predictions, we calculate the force again
using both direct pressure integration and the added-mass X
3
Z=H…t† ˆ An W 3⫺2n : …62†
formula. The vertical added mass of the immersed section of nˆ1
the circular cylinder is evaluated using the formula from
Taylor [18]: The velocity potential f can now be written in the form

M a ˆ C a rY 2 ; with f…y 0 ; z 0 ; t† ˆ Re{ ⫺ iA1 VH…W ⫺ W ⫺1 †}; …63†


and the complex velocity is given by
Ca ˆ p…v2 ⫹ 2†=3 ⫺ 2‰p…1 ⫺ 1=v†csc2 …p=v† ⫹ tan⫺1 …p=v†Š: " #⫺1
…57† ⫺2
X3
u ⫺ iv ˆ ⫺iA1 V…1 ⫹ W † …3 ⫺ 2n†An W 2⫺2n
: …64†
The comparisons are shown in Fig. 11, where the results nˆ1

obtained using the theories of von Kármán [3] and Wagner Givenf for any specified Y, the evolution of the intersection
[1] are also plotted. The present theoretical prediction based point is determined by following the general procedure in
on direct pressure integration agrees very well with the Section 3.2.
experiments. The small discrepancies can likely be attribu- To illustrate the accuracy and usefulness of this scheme,
ted to three-dimensional effects in the experiments. As in we apply (58) and (62) to the wedge and circle we consid-
the case of the wedge, the prediction using added mass ered in Sections 4 and 5. For the wedge, L ˆ cot a; s ˆ 1=2;
slightly overestimates the impact force. Again, as expected, and the coefficients A1 ; A2 and A3 are functions of a , but
Wagner’s theory gives a significantly higher force while von independent of t according to (59)–(61). From Eq. (64), the
Kármán’s theory under predicts the force coefficient. velocity of the free surface v0 …Y…t†; `…t†† is obtained, which
is then used to determine the influence coefficient bn in Eq.
6. Application to general ship sections (22). From Eq. (21), the unknown coefficients an are solved.
We obtain finally the solution for the splash-up coefficient
For general body geometries, a closed-form solution to for any deadrise angle a
the ‘‘double-body’’ boundary-value problem for f may not Z∞ A1 …1 ⫹ p⫺2 † cot a
be available. For ship-like sections, however, a semi-analy- g… a † ˆ dp: …65†
1 …A1 p ⫹ A2 p⫺1 ⫹ A3 p⫺3 †2
tical closed-form solution can be obtained by making use of
a Lewis-form approximation (e.g. [9,10]) to the body This solution, derived using the Lewis-form approximation,
surface. is plotted in Fig. 3. The agreement with the exact solution of
To solve for f at any instant, we first approximate the Section 4.1 is excellent.
14 X. Mei et al. / Applied Ocean Research 21 (1999) 1–15

Fig. 12. Comparison between the present theory (—) and the measurement of Zhao, Faltinsen and Aarsnes [7] (– – –) for the time history of the impact force
on a bow-flare section free-falling into water.

For the case of a circular cylinder, we have L ˆ Y= wedges of arbitrary deadrise angles and circular cylinder,
‰1 ⫺ …1 ⫺ Y 2 †1=2 Š and s ˆ ‰arcsin Y ⫺ Y…1 ⫺ Y 2 †1=2 Š= our analytic solutions compare well with measurements and
‰Y ⫺ Y…1 ⫺ Y † Š: With Eq. (64), the influence coeffi-
2 1=2
with a fully nonlinear numerical simulation using a spray-jet
cients bn can be evaluated from Eq. (22), and the coefficients matching scheme, which we developed. Comparisons are
an are solved from Eq. (21). The result for the evolution of the also made to existing asymptotic theories and approxima-
intersection point, obtained with N ˆ 10, is shown in Fig. tions. For application to general ship-like body sections, we
10. The result obtained with the Lewis-form approximation obtain the solution to the impact of a Lewis-form body. This
compares very well with the exact solution of Section 5. is applied to a bow-flare section and the result compares
As a final application, we apply this analysis to the bow- favorable to measurement. The generalization of Wagner’s
flare section used by Zhao et al. [7] in their experiment. The theoretical solution greatly increases its validity in practical
section geometry, given in Zhao et al. [7], has a half width of applications and eliminates the need for nonlinear numerical
B ˆ 0:155 m and the vertical distance from keel to knuckle simulations in the impact study of many classes of body
is H ˆ 0:203 m. In the tests, the section is in free fall so that sections.
the downward body velocity changes slightly with time
during the short duration of the impact measurement.
From their measurement (their Fig. 18), we estimate the Acknowledgements
average vertical velocity to be approximately 2.6 m/s. In
our analysis, we approximate the section geometry by This research is financially support by the Office of Naval
Lewis form and, for simplicity, use a constant value of Research.
V ˆ 2:6 m=s. The comparison for the time history of the
impact force between the present theory and measurement
of Zhao et al. [7] is given in Fig. 12.
References

[1] Wagner H. Uber Stoss-und Gleitvorgange an der Oberflache von


7. Conclusions Flussigkeiten. Z. angew. Math. Mech. 1932;12(4):193–215.
[2] Korobkin AA, Pukhnachov VV. Initial stage of water impact. Ann.
We develop an analytic solution for the water impact Rev. Fluid Mech. 1988;20:159–185.
problem of general two-dimensional bodies. The approach [3] Von Karman T. The impact of seaplane floats during landing. NACA
TN, 1929.
can be considered as a generalization of the method of [4] Armand JL, Cointe, R. Hydrodynamic impact analysis of a cylinder.
Wagner [1] but with the body boundary condition now satis- Proc. Fifth Inter. Offshore Mech. and Arctic Engng. Symp., Tokyo,
fied on the exact instantaneous position of the body. For Japan, vol. 1. 1987:609–634.
X. Mei et al. / Applied Ocean Research 21 (1999) 1–15 15

[5] Howison SD, Ockendon JR, Wilson SK. Incompressible water-entry [13] Faltinsen O. Sea loads on ships and offshore structures. Cambridge:
problems at small deadrise angles. J. Fluid Mech. 1991;222:215–230. University Press, 1990.
[6] Dobrovol’skaya ZN. On some problems of similarity flow of fluids [14] Lin WM. Nonlinear motion of the free surface near a moving body.
with a free surface. J. Fluid Mech. 1969;36:805–829. Ph.D Thesis, MIT, 1984.
[7] Zhao R, Faltinsen O, Aarsnes J. Water entry of arbitrary two-dimen- [15] Greenhow M. Wedge entry into initially calm water. Appl. Ocean
sional sections with and without flow separation. ONR, Norway, 1996. Res. 1987;9:214–223.
[8] Zhao R, Faltinsen O. Water entry of two-dimensional bodies. J. Fluid [16] Dommermuth DG, Yue DKP, Lin WM, Rapp RJ, Chan ES,
Mech. 1993;246:593–612. Melville WK. Deep-water plunging breakers:, a comparison
[9] Landweber L, Macagno M. Added masses of two-dimensional forms between potential theory and experiments. J. Fluid Mech. 1988;189:
by conformal mapping. J. Ship Res. 1967;11:109–116. 423–442.
[10] Kerczek CV, Tuck EO. The representation of ship hulls by conformal [17] Campbell IMC, Weynberg PA. Measurement of parameters affecting
mapping functions. J. Ship Res. 1969; 284–298. slamming. Final Report, Rep. No. 440, Technology Reports Center
[11] Vinje T, Brevig P. Nonlinear two dimensional ship motions. Report No. OT-R-8042. Southampton Univeristy: Wolfson Unit for Marine
R-112.81, The Ship Research Institute of Norway, 1981. Technology, 1980.
[12] Payne PR. Recent developments in ‘‘added-mass’’ planing theory. [18] Taylor JL. Hydrodynamical inertia coefficients. Philosophical Maga-
Ocean Engng. 1994;21(3):257–309. zine 1930;9:161–183.

You might also like